You are on page 1of 16

Journal of Experimental Botany, Vol. 70, No. 16 pp.

4139–4154, 2019
doi:10.1093/jxb/erz213  Advance Access Publication 4 May, 2019

REVIEW PAPER

Rhodanese domain-containing sulfurtransferases:


multifaceted proteins involved in sulfur trafficking in plants
Benjamin Selles, Anna Moseler, Nicolas Rouhier and Jérémy Couturier*,
Université de Lorraine, Inra, IAM, F-54000 Nancy, France

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


*  Correspondence: jeremy.couturier@univ-lorraine.fr

Received 17 April 2019; Editorial decision 29 April 2019; Accepted 29 April 2019

Editor: Stanislav Kopriva, University of Cologne, Germany

Abstract
Sulfur is an essential element for the growth and development of plants, which synthesize cysteine and methionine
from the reductive assimilation of sulfate. Besides its incorporation into proteins, cysteine is the building block for the
biosynthesis of numerous sulfur-containing molecules and cofactors. The required sulfur atoms are extracted either
directly from cysteine by cysteine desulfurases or indirectly after its catabolic transformation to 3-mercaptopyruvate,
a substrate for sulfurtransferases (STRs). Both enzymes are transiently persulfidated in their reaction cycle, i.e. the
abstracted sulfur atom is bound to a reactive cysteine residue in the form of a persulfide group. Trans-persulfidation
reactions occur when sulfur atoms are transferred to nucleophilic acceptors such as glutathione, proteins, or small
metabolites. STRs form a ubiquitous, multigenic protein family. They are characterized by the presence of at least
one rhodanese homology domain (Rhd), which usually contains the catalytic, persulfidated cysteine. In this review,
we focus on Arabidopsis STRs, presenting the sequence characteristics of all family members as well as their bio-
chemical and structural features. The physiological functions of particular STRs in the biosynthesis of molybdenum
cofactor, thio-modification of cytosolic tRNAs, arsenate tolerance, cysteine catabolism, and hydrogen sulfide forma-
tion are also discussed.

Keywords:   Cysteine, hydrogen sulfide signaling, persulfide group, rhodanese, sulfur trafficking, sulfurtransferase.

Introduction
Unlike animals, which cannot assimilate inorganic sulfur and action of glutathione transferase (Su et al., 2011). In the case of
thus need to absorb cysteine or methionine, plants achieve the thionucleoside formation or Fe–S cluster and Moco synthesis,
reductive assimilation of sulfate to synthesize cysteine, forming other sulfur sources are used (Mueller, 2006). For instance, the
sulfite and sulfide as plastidial intermediates, and subsequently pyridoxal 5′-phosphate-dependent cysteine desulfurases (CDs)
methionine and glutathione (GSH) (Takahashi et  al., 2011). catalyse the desulfuration of cysteine (Zheng et  al., 1993),
Other key cellular components, such as vitamins (biotin, thia- leading to the formation of a persulfide enzyme intermediate
mine, lipoic acid), tRNAs containing thionucleosides, or co- and the concomitant release of alanine (Behshad and Bollinger,
factors such as iron–sulfur (Fe–S) clusters and molybdenum 2009). The efficiency and specificity of sulfur transfer to ac-
cofactor (Moco), contain sulfur atoms coming from various ceptor molecules vary according to the subclass of CDs and
origins. In the case of vitamin or camalexin biosynthesis, sulfur the type of sulfur acceptors, the nature of which dictates the
is respectively provided by the degradation of Fe–S clus- direction and flow of the sulfur transfer reaction. Whereas
ters (Lanz and Booker, 2015) or by a GSH molecule via the glutathione persulfide (GSSH) may represent a relatively stable

© The Author(s) 2019. Published by Oxford University Press on behalf of the Society for Experimental Biology. All rights reserved.
For permissions, please email: journals.permissions@oup.com
4140 | Selles et al.

and specific transport form of sulfur atoms (Hildebrandt and organisms whereas other clusters are specific (clusters III and
Grieshaber, 2008; Ida et al., 2014), carrier proteins ensure the V) or contain an increased number of representatives in plants
specific transfer of sulfur to acceptors. This function is pri- (clusters II, IV and VI). In the context of this review on the in-
marily achieved by sulfurtransferases (STRs; E.C.2.8.1.x). volvement of plant STRs in sulfur transfer reactions, those that
They possess one or several rhodanese (Rhd) domains usu- belong to cluster II and do not possess the conserved catalytic
ally including a reactive catalytic cysteine (Hatzfeld and Saito, cysteine in the Rhd domain will not be described in detail.
2000; Bordo and Bork, 2002; Cipollone et al., 2007; Guretzki Hence, we discuss the recent genetic, biochemical, structural,
and Papenbrock, 2011). Hence, most STRs catalyse the transfer and molecular studies that have contributed to improving our
of a sulfur atom from donors to nucleophilic acceptors by understanding of the physiological roles of STRs in plants.
forming intermediate protein persulfides on a reactive cata- After illustrating the diversity of plant STRs in terms of pri-
lytic cysteine.When two polypeptides are involved, the transfer mary sequences, protein domain architecture, subcellular lo-
of persulfide groups is referred to as a trans-persulfidation calization and gene expression using Arabidopsis isoforms as
reaction. Although STRs are ubiquitously distributed, they representatives, we focus our attention on the documented
differ quite considerably in terms of primary sequences, pro- roles of STRs in molybdopterin synthesis, thionucleoside

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


tein domain architecture, and structure/length of the active modification, arsenate tolerance, cysteine catabolism, and the
site loop (Bordo and Bork, 2002). Nonetheless, three major biosynthesis of hydrogen sulfide (H2S), a signaling molecule
groups can be differentiated (Bordo and Bork, 2002; Cipollone promoting protein persulfidation.
et  al., 2007; Bartels et  al., 2007). The first one comprises
STRs formed by a single Rhd domain that include so-called
CDC25 phosphatases and thiosulfate sulfurtransferases (TSTs; Diversity of plant STR isoforms: focus on
EC.2.8.1.1) according to their ability to use thiosulfate as a Arabidopsis
sulfur donor in vitro. The presence of an additional residue in
the active-site loop of CDC25 phosphatases (seven residues The Arabidopsis nuclear genome was initially reported to en-
compared with six in TSTs) constitutes an important differ- code 20 STR sequences (Papenbrock et al., 2011). However, a
ence involved in the capacity of CDC25 to dephosphorylate recent careful analysis led us to include an additional protein
tyrosine residues (Bordo and Bork, 2002). The second major containing a Rhd domain, which was numbered consecutively
group is formed by STRs containing two Rhd domains as AtSTR19 (At3g59780), even though it does not have the
but only the C-terminal one possesses the catalytic cyst- conserved cysteine residue (Moseler et al., 2019).
eine. They are present in most organisms. According to their
ability to use 3-mercaptopyruvate (3-MP) as a substrate in Domain organization, expression, and subcellular
vitro, they are referred to as 3-MP-STRs or mercaptopyruvate localization
sulfurtransferases (MSTs; EC.2.8.1.2) (Nakamura et al., 2000;
Papenbrock and Schmidt, 2000; Colnaghi et  al., 2001; Yadav According to the variability observed among STRs in terms
et  al., 2013). It is worth mentioning that bacteria and mam- of primary sequence and domain organization (including po-
mals possess additional STR isoforms with two Rhod domains tential transmembrane spans), the Arabidopsis STR equipment
(Bordo and Bork, 2002; Nambi et al., 2015), such as the bo- forms nine different clusters (Table 1; Moseler et  al., 2019).
vine rhodanese protein named Rhobov, which display TST STRs composed of two Rhd domains form the first group
activity (Blumenthal and Heinrikson, 1971). The mammalian comprising so-called MSTs, i.e. STR1 and STR2, which
rhodanese isoforms have a CRXGX[R/T] active site signa- are 379 and 342 residues long, respectively (Table 1). STR3
ture instead of the characteristic CG[S/T]GVT motif found in (387 residues), STR4 (466 residues), STR4a (264 residues),
MSTs (Bordo and Bork, 2002; Cipollone et al., 2007). In fact, and STR19 (686 residues) form cluster II as they contain a
interchanging the two residues following the catalytic cysteine single Rhd domain associated with one, two, one, and three
of rat rhodanese and MST by site-directed mutagenesis was predicted transmembrane regions, respectively (Table 1). Note
sufficient to change the activity profile from one to the other that STR3, STR4, and STR19 are considered as inactive
(Nagahara et  al., 1995; Nagahara and Nishino, 1996). Finally, STRs because the catalytic cysteine is replaced by an aspar-
the third main group is formed by STRs containing a Rhd do- tate. Cluster III contains the sole STR5 isoform. It possesses
main possessing the catalytic cysteine fused to a domain having a single Rhd domain exhibiting the catalytic cysteine but in a
another function conferring them specific roles (Bordo and quite divergent HCALSQVR signature typical of the general
Bork, 2002; Cipollone et al., 2007). HCX5R signature of CDC25 phosphatases. STR6, STR7, and
The recent comparative analysis of 25 genomes of photosyn- STR8 belong to cluster IV. They have also a single Rhd do-
thetic organisms has provided a detailed classification and in- main with a strictly conserved CTGGIR signature, but they
formation about evolutionary features of STRs pointing to the are elongated (581, 474, and 448 residues respectively) on both
expansion of this family in higher plants. Hence, in photosyn- sides of the Rhd domain.The next group (cluster V) comprises
thetic organisms, the STR family comprises 5–35 genes, and STR9, STR10, and STR11 isoforms (214–292 residues) cor-
was divided into nine clusters according to the STR primary responding to proteins with a single Rhd domain exhibiting a
sequences and domain arrangements (Moseler et  al., 2019). catalytic cysteine and a predicted C-terminal transmembrane
Plant STR isoforms of clusters I  (MSTs), VII (multidomain segment (Table 1). STR12 (299 residues) and STR13 (464 res-
STRs), VIII (TSTs) and IX (TSTs) possess orthologs in most idues) form two additional groups (clusters VI and VII), having
Roles of sulfurtransferases in sulfur trafficking in plants  |  4141

Table 1.  Subcellular localization and domain organization of the STR/rhodanese family members from Arabidopsis

Name (ac- Cluster Length Active Domain arrangement Subcellular localization


cession no.) (aa) site motif
STR1 I 379 CGTGVT Mitochondrion
(At1g79230) GFP fusion (Hatzfeld and Saito, 2000; Nakamura et al.,
2000; Bauer et al., 2004) and immunoblot analysis
(Nakamura et al., 2000; Bauer et al., 2004)
STR2 I 342 CGTGVT Cytosol
(At1g16460) GFP fusion (Hatzfeld and Saito, 2000; Nakamura et al.,
2000; Bauer et al., 2004) and immunoblot analysis
(Nakamura et al., 2000; Bauer et al., 2004)
STR3 II 387 DSYTDS Chloroplast (thylakoid membrane)
(At5g23060) GFP/YFP fusion (Weinl et al., 2008; Nomura et al.,
2008; Vainonen et al., 2008), immunoblot analysis

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


(Weinl et al., 2008; Vainonen et al., 2008) and prote-
omics (Peltier et al., 2004)
STR4/TROL II 466 DKFDGN Chloroplast (inner membrane envelope and thyla-
(At4g01050) koid)
YFP fusion (Jurić et al., 2009) and proteomics (Friso
et al., 2004; Peltier et al., 2004)
STR4a II 264 CVLDNF Chloroplast (thylakoid)
(At3g25480) Proteomics (Peltier et al., 2004; Tomizioli et al., 2014)
STR5/ III 146 CALSQVR Mitochondrion
CDC25/ACR2 GFP fusion (Narsai et al., 2011)
(At5g03455)
STR6 IV 581 CTGGIR Unknown
(At1g09280)
STR7 IV 474 CTGGIR Unknown
(At2g40760)
STR8 IV 448 CTGGIR Chloroplast (thylakoid membrane)
(At1g17850) Proteomics (Tomizioli et al., 2014)
STR9 V 234 CQEGLR Chloroplast (thylakoid membrane)
(At2g42220) Proteomics (Tomizioli et al., 2014)
STR10 V 214 CGEGLR Mitochondrion
(At3g08920) GFP fusion (Bartels, 2006)
STR11 V 292 CQKGLR Chloroplast (thylakoid membrane)
(At4g24750) GFP fusion (Bartels, 2006) and proteomics (Tomizioli
et al., 2014)
STR12 VI 299 CKVGGR Chloroplast (stroma)
(At5g19370) Proteomics (Tomizioli et al., 2014)
STR13/CNX5 VII 464 CRRGND Cytosol
(At5g55130) Bimolecular fluorescence complementation (Kaufholdt
et al., 2013)
STR14 VIII 224 CSSAGT Chloroplast (envelope, thylakoid)
(At4g27700) GFP fusion (Bauer et al., 2004)
STR15/Sen1 IX 182 CESGQM Chloroplast (thylakoid membrane)
(At4g35770) GFP fusion (Bauer et al., 2004) and electronic micros-
copy (Bauer et al., 2004)
STR16 IX 120 CQSGGR Chloroplast
(At5g66040) GFP fusion (Bauer et al., 2004)
STR17 IX 156 CKSGVR Unknown
(At2g17850)
STR17a/ IX 169 CNAGGR Cytosol/nucleus
ARQ1/HAC1 GFP fusion (Sánchez-Bermejo et al., 2014)
(At2g21045)
STR18 IX 136 CQSGAR Cytosol
(At5g66170) GFP fusion (Bauer et al., 2004)
STR19 II 686 DADGTR Chloroplast
(At3g59780) Proteomics (Peltier et al., 2004; Tomizioli et al., 2014)

The catalytic cysteine is indicated in bold. Domain prediction is based on Pfam (http://pfam.xfam.org/) and Interpro (http://www.ebi.ac.uk/interpro/).
FSH1, serine hydrolase; MoeB, molybdopterin biosynthesis; Rota, rotamase.
4142 | Selles et al.

an N-terminal rotamase and MoeB domain, respectively, in been demonstrated or predicted as targeted to the chloroplasts/
addition to the active Rhd domain (Table 1). Cluster VIII con- plastids (Bauer et al., 2004; Friso et al., 2004; Peltier et al., 2004;
tains only AtSTR14 because the Rhd domain has a specific Bartels, 2006; Nomura et al., 2008; Vainonen et al., 2008; Weinl
CSSAGT signature and it displays a short N-terminal extension et al., 2008; Jurić et al., 2009; Tomizioli et al., 2014) are encoded
forming a protein of 224 residues. Finally, cluster IX comprises by genes present in these two clusters (Table 1; Fig. 1). With re-
STR15–STR18. They are all short STR versions of less than gards to genes present in the third cluster (STR1, STR6, STR13,
182 amino acids, the difference residing in the active site sig- STR17, STR17a, and STR18), their expression does not vary
nature and in the presence of N-terminal targeting sequences, much among organs but they have in common a highest tran-
but none has additional recognizable domains, except a pre- script level in dry and imbibed seeds and in roots. Note that
dicted C-terminal transmembrane segment in STR15 likely AtSTR17a is specifically more expressed in roots. The proteins
promoting its localization to thylakoids (Bauer et al., 2004). encoded by these genes are expressed in the cytosol (STR13,
As an additional criterion to understand the diversity and evo- STR17a, and STR18) (Bauer et  al., 2004), in mitochondria
lution of the STR family in plants, the gene expression pattern (STR1) (Hatzfeld and Saito, 2000; Nakamura et al., 2000; Bauer
and protein subcellular localization of Arabidopsis STRs was et al., 2004), or have unknown localizations (STR6 and STR17)

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


analysed. A recent, high resolution developmental transcript pro- having no typical targeting sequences. These diverse subcellular
file was analysed (Winter et al., 2007; Klepikova et al., 2016) and localizations and gene expression patterns are understandable
selected tissues and stages were represented as a heat map (Fig. considering the existence of numerous sulfur-containing com-
1). Three main clusters are distinguished. A first cluster regroups pounds in the different subcellular compartments of plant cells
STR2, STR5, STR15, and STR16 representatives. They are and thus the need of specific systems to mobilize and transfer
mostly expressed in leaves and hypocotyl/stems (Fig. 1). STRs sulfur for their biosynthesis or modification.
present in the second cluster (STR3, STR4, STR4a, STR7, STR8,
STR9, STR10, STR11, STR12, and STR14) have in common a Biochemical and structural properties
higher expression level in cotyledons, mature leaves (8–12) and
vegetative rosette before or after the transition to flowering (Fig. Despite the described diversity of STR primary
1). Two sub-clusters can be distinguished mainly based on their sequences, the fold adopted by the Rhd domain consists
differential expression in seeds. Consistently, all STRs that have of a β1α1β2α2β3α3β4α4β5α5 topology organizing into a

STR16
STR5
STR2
STR15
STR11
STR4a
STR12
STR7
STR8
STR14
STR4
STR9
STR10
STR3
STR18
STR17
STR13
STR1
STR6
STR17a

Fig. 1.  Heatmap representation of STR transcript profiles at selected developmental stages and tissues of Arabidopsis. Gene expression data for each
Arabidopsis STR were retrieved from the ‘Developmental Map’ tab of the Arabidopsis eFP Browser database (http://bar.utoronto.ca/efp/cgi-bin/efpWeb.
cgi) (Winter et al., 2007; Klepikova et al., 2016). No data were available for STR19. Numerical data were extracted and represented as a heatmap
using the expression tool of Heatmapper software (http://www2.heatmapper.ca/expression/) (Babicki et al., 2016). The parameters applied are score
representation by row, clustering method by average linkage, and Pearson distance measurement.
Roles of sulfurtransferases in sulfur trafficking in plants  |  4143

well-conserved central β-sheet composed of five strands sur- otherwise indicated), is highly conserved in Arabidopsis
rounded by four to five α-helices (Ploegman et al., 1978). Up paralogs (Fig. 2A). It is also conserved in representatives of
to now, three-dimensional structures have been solved for the STR5/CDC25 family and more generally in the Rhd
AtSTR5/CDC25, AtSTR16, and the inactive Rhd domain of domains of all plant STRs (Moseler et al., 2019). The residues
AtSTR4/TROL (Landrieu et al., 2004a; Pantoja-Uceda et al., are located at the end of β2 (Fig. 2B) and were proposed
2005; Cornilescu et al., 2006). A structure-based amino acid to participate in AtSTR16 stability (Cornilescu et al., 2006).
alignment of Arabidopsis STRs possessing the catalytic cyst- The catalytic cysteine (Cys80) is strictly conserved and is pre-
eine, using AtSTR16 NMR structure (PDB 1TQ1) as a tem- sent in the loop between β4 and α4 secondary structures and
plate, indicates that only a few residues are strictly conserved forms with the five or six (in the case of CDC25) following
(Fig. 2A). In the N-terminal part of the Rhd domain, a three amino acids the so-called active site loop of Rhd domain (Figs
residue motif, D28[V/A]29R30 (AtSTR16 numbering unless 2B, 3). The residues forming the active site are quite variable.

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023

Fig. 2.  Structure-based sequence alignment of Rhd domains of Arabidopsis STR isoforms. The sequences corresponding to the Rhd domain of
Arabidopsis STRs possessing a cysteine residue were aligned with MUSCLE (https://www.ebi.ac.uk/Tools/msa/muscle/). Colored alignment (A) and the
corresponding structure representing the structural conservation level at each amino acid position (B) were generated by CONSURF (http://consurf.tau.
ac.il/2016/) using the PDB file of the AtSTR16 NMR structure (PDB 1TQ1). The sequence numbering and secondary structures correspond to AtSTR16.
In (A) the catalytic cysteine residue is marked by an asterisk and amino acids studied by mutagenesis indicated by triangles. In (B) the residues studied
by mutagenesis in various selected STR representatives are shown with their lateral chains as stick and labeled. The most conserved structural positions
(D28[V/A]29R30, C80, G82, L93, G97, and G107) are also labeled whereas green parts correspond to regions with insufficient data to represent a conservation.
4144 | Selles et al.

Val251
Gly250 Val252
Gly251
Rhobov / TST Thr252 Thr253 Human 3-MST / MST
Lys249
CRKGVT Ser250 CGSGVT
Cys247
Cys248
Arg248 Gly249

Arg85 Lys83
AtSTR16 / TST Gly82 Human TSTD / TST
Gly83
Ser82 Gly84 Met81
CQSGGR CQMGKR
Arg84

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


Cys79
Gln81 Cys80 Gln80

Fig. 3.  Structural comparison of the active site loops of TST-type and MST-type STRs. The representation of the six amino acids forming the so-called
active site loop from bovine Rhobov (PDB 1RHD), human 3-MST (PDB 4JGT), Arabidopsis STR16 (PDB 1TQ1), and human TSTD (PDB 6BEV) was
generated using PyMOL.

A characteristic signature was defined for MSTs (CGSGVT) As the biochemical data concerning plant STRs are rela-
and TSTs (CRKGVT) and it is likely that this relates at least tively scarce, we consider here mutagenesis studies conducted
in part to electrostatic differences of their substrates, as ex- on non-plant STR orthologs such as bovine rhodanese
emplified by 3-MP and thiosulfate, respectively (Bordo and (Rhobov), the Escherichia coli MST ortholog SseA, the
Bork, 2002). At positions +1 and +2 after the catalytic cyst- Azotobacter vinelandii RhdA protein and the Rhd domain of
eine, bulky and charged residues (Arg/Lys) are found in TSTs human MOCS3 to identify other residues potentially im-
whereas smaller and uncharged residues (Gly/Ser) are pre- portant for the structure and activity of plant STRs (Table 2).
sent in MSTs (Fig. 3). The Gly residue usually at position +3 Structural analysis of human MST and biochemical charac-
of the catalytic cysteine and the Leu residue at position 93 terization of E. coli SseA indicated that two arginine residues,
(Fig. 2A) are conserved in all Arabidopsis STRs even though Arg178 and Arg187 (position 34 and insertion between positions
the AtSTR5 signature containing the catalytic cysteine does 36 and 37, respectively), play a role in substrate recognition,
not align at all. Two other Gly residues at positions 97 and being involved in the proper positioning of the substrate rela-
107 are present in all plant STR isoforms except in STR1/2 tive to the catalytic Cys237 (Figs 2A, 3) (Yadav et al., 2013; Lec
orthologs in which Gly107 is replaced by a serine. These three et  al., 2018). Hence in EcSseA, the change of Arg178 to Leu
Gly residues are located just before and at the end of α4, leads to a decrease of protein affinity towards 3-MP (Lec et al.,
and at the initiation of α5, respectively (Fig. 2B). Despite the 2018). Both AtSTR1 and AtSTR2 (and AtSTR5) possess this
high conservation of these four residues (Gly83, Leu93, Gly97, Arg residue whereas most other STRs possess a Glu residue
Gly107), their roles in STR structure/function are unknown. (Fig. 2A). On the contrary, Arg187 is only present in AtSTR1
On the contrary, the importance of other residues was inves- and is located into an insertion sequence specific to MST
tigated more deeply using both plant and non-plant STRs orthologs (Fig. 2). Interestingly, the mutation of Arg187, which
(Table 2). Hence, besides the mandatory catalytic cysteine is part of the hydrophobic pocket of the active site (Ploegman
(Cys80 in Fig. 2), other residues are important for the catalytic et al., 1978), into a Leu in bovine Rhobov, a TST-type STR,
activity of the Rhd domain. For example, the change of the leads to a 20-fold decrease of affinity towards thiosulfate (Km
cysteine residue close to the active site in AtSTR1 (Cys339, shifted from 3.7 to 73.2 mM in the R187L variant) (Luo and
position 87) to a valine residue led to a 4-fold decrease of the Horowitz, 1994). Hence, this residue may also be important for
TST activity whereas the MST activity was slightly reduced thiosulfate accommodation in the active site.
(Burow et al., 2002). However, although highly conserved in It is expected that other residues found in the active site
MSTs from terrestrial plants, this cysteine is not present in loops of MST, TST, or CDC25 generate the necessary se-
other STRs. For instance, it corresponds to Ile87 of AtSTR16, lectivity of these proteins towards their respective substrates
a position presenting a weak conservation score (Fig. 2A). (Bordo and Bork, 2002; Cipollone et al., 2007; Lec et al., 2018;
Substitutions of the two other Cys residues, Cys294 to Asn Libiad et al., 2018).Thus, apart from the Arg residues described
and Cys304 to Asp, found in the active C-terminal Rhd do- above, the importance of the residues present at positions +2
main of AtSTR1 had minor effects with a slight decrease and +5 of the catalytic Cys in regular STRs was investigated.
of both Km and kcat towards thiosulfate (Burow et al., 2002). For MST-type STR proteins, studies performed using
According to the alignment, the corresponding amino acids EcSseA, human and Leishmania major MST variants initially
in AtSTR16 (positions 43 and 53) are found respectively in suggested that the +2 residue (Ser82 in Figs 2A, 3) is involved in
the α2–β3 and β3–α3 loops and exhibit a low conservation the deprotonation of 3-MP and in the stabilization of the pyru-
score (Fig. 2B). vate enolate (Colnaghi et al., 2001; Alphey et al., 2003; Huang
Roles of sulfurtransferases in sulfur trafficking in plants  |  4145

Table 2.  Mutagenesis studies performed using STR isoforms

Protein Uniprot Ref Type Position in Corresponding Substi- Catalytic effect Reference
name the original position in tution
sequence AtSTR16
Rhobov P00586 TST 187 34 R→L Decreased reactivity to- Luo and Horowitz
wards thiosulfate (1994)
SseA P31142 MST 178 34 R→L Decreased reactivity to- Lec et al. (2018)
wards 3-MP
STR1 O64530 MST 294 43 C→N Slightly reduced TST ac- Burow et al. (2002)
tivity, no impact on MST
activity
STR1 O64530 MST 304 53 C→E Slightly reduced TST ac- Burow et al. (2002)
tivity, no impact on MST
activity

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


STR1 O64530 MST 333 80 C→S Abolished TST and MST Burow et al. (2002)
activities
MOCS3 O95396 — 412 80 C→A Abolished STR activity Matthies et al.
(2004)
RhdA P52197 TST 232 82 T→K Increased TST activity Pagani et al. (2000)
RhdA P52198 TST 232 82 T→A Increased TST activity Pagani et al. (2000)
Rhobov P00586 TST 250 82 K→A Abolished TST activity Luo and Horowitz
(1994)
SseA P31142 MST 239 82 S→A Decreased affinity for Colnaghi et al.
3-MP; abolished thiosul- (2001), Lec et al.
fate binding (2018)
SseA P31142 MST 239 82 S→K Decreased MST activity; Colnaghi et al.
increased affinity for thio- (2001)
sulfate
MOCS3 O95396 — 417 85 D→R 470-fold increased activity Krepinsky and
Leimkühler (2007)
MOCS3 O95396 — 417 85 D→T 90-fold increased activity Krepinsky and
Leimkühler (2007)
STR1 O64530 MST 339 87 C→V 4-fold decreased TST Burow et al. (2002)
activity; slightly reduced
MST activity

Summary table describing the catalytic effects of STR residue mutations. The nature and position of the mutations are indicated as well as the
corresponding AtSTR16 numbering used in Fig. 2. The catalytic cysteine is indicated in bold.

and Yu, 2016). However, another study, focusing on catalytic Finally, the importance of the residue at position +5 (Arg85
cysteine persulfide formation and regeneration by thioredoxin in Fig. 2) for protein activity was also investigated. Most STRs
(TRX) of EcSseA and human MST, reported different results. have an Arg or sometimes a Thr but AtSTR13 and human
In this work, it was found that the step of sulfur transfer be- MOCS3 have an Asp, which makes a considerable difference.
tween 3-MP and the thiol group of the catalytic cysteine, and In human MOCS3, the change of this Asp417 to an Arg or a
the protonation state of 3-MP at various pH values is not af- Thr results in a strong increase of protein activity towards thio-
fected in the corresponding S239A variant (Lec et al., 2018). sulfate (Table 2; Krepinsky and Leimkühler, 2007). Hence, the
These contradictory results raise questions about the import- fact that AtSTR13 and MOCS3 possess an N-terminal MoeB
ance of the residue found at the +2 position within the active domain required for Moco synthesis and thus have particular
site loop of MST-type STRs. substrates/partners may explain why the basic residue usually
Regarding TST-type proteins, the residue at this position is found at this position and which seems important for the pro-
important for their activity but again effects in protein variants tein activity with thiosulfate has been replaced by an acidic
are variable and sometimes opposite. The change of Lys250 to residue.
Ala in bovine Rhobov abolishes its activity towards thiosul-
fate whereas it does not affect activity towards more complex
substrates such as thiosulfonate (Table 2; Luo and Horowitz, Dual functions of STR13 in the delivery of
1994), pointing to substrate-dependent effects. A  mutational sulfur atoms
analysis was also performed using A. vinelandii RhdA, an atyp- Implication in Moco synthesis
ical enzyme that uses only thiosulfate as a sulfur donor in vitro.
Contrary to expectation, the change of Thr232 to Lys or Ala Moco consists of a molybdenum atom covalently bound
leads in fact to an increase of TST activity (Pagani et al., 2000). to two sulfur atoms of a tricyclic pterin moiety referred to
4146 | Selles et al.

as molybdopterin (MPT). Moco is found at the active site Moco into apoproteins. In NR and SO, a sulfur atom of a
of molybdenum enzymes. In plants, the Moco-dependent conserved protein cysteine residue covalently binds the Mo
enzymes nitrate reductase (NR), aldehyde oxidase (AO), atom. On the contrary in XDH and AO, the Moco sulfur-
xanthine dehydrogenase (XDH), and sulfite oxidase (SO) ation is independent of protein cysteine residues and de-
function respectively in nitrogen assimilation, abscisic acid pends on the activity of the cytosolic cysteine desulfurase
synthesis, purine catabolism, and sulfite detoxification ABA3 (Bittner et al., 2001). This CD is relatively atypical as
(Schwarz and Mendel, 2006). Moco biosynthesis is there- it consists of two domains, an N-terminal CD domain that
fore crucial and is now well understood (Fig. 4A). The first transfers the sulfur atom to a Moco bound to the C-terminal
step, consisting of the condensation of GTP into cyclic domain, prior to the transfer and insertion of this sulfurated
pyranopterin monophosphate (cPMP), is catalysed by the Moco into AO and XDH (Wollers et al., 2008). A functional
Fe–S protein CNX2 and by CNX3 and takes place in the characterization demonstrated that an Arabidopsis cnx5/str13
mitochondrial matrix. cPMP is then exported to the cytosol. mutant containing a change of the conserved Ser149 residue
Initially it was suggested that the ABC transporter ATM3 to Phe accumulates high levels of cPMP, whereas cnx5 null
transports cPMP, as the corresponding mutants show a mutants have a strong growth defect and are sterile (Nakai

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


strong phenotype for Moco-containing proteins such as NR et al., 2012; Kruse et al., 2018). STR13, also named CNX5/
(Bernard et al., 2009; Teschner et al., 2010). Nevertheless, it MOCS3, has the particularity of containing a MoeB-like
was shown that atm3 mutants contain a lower content of domain necessary for MPT biosynthesis in the N-terminal
cPMP but a more oxidized mitochondrial matrix, suggesting region associated to a Rhd domain with a CRRGND ac-
that the role of ATM3 in the Moco biosynthesis pathway tive site motif. In vitro assays performed with the human
may be indirect (Kruse et al., 2018). In the cytosol, two sulfur ortholog, MOCS3, revealed that the persulfidated Rhd do-
atoms are inserted into cPMP by the concerted action of main of MOCS3 is able to provide sulfur to MOCS2A, the
CNX5, CNX6, and CNX7, resulting in the formation of human ortholog of CNX7, enabling the synthesis of MPT
MPT (Kaufholdt et al., 2013). Thereafter, CNX1 inserts the (Matthies et al., 2004). Moreover, mutating the catalytic cyst-
molybdenum atom, which is coordinated by a dithiolene eine of the Rhd domain prevents the formation of MPT
group, leading to the formation of Moco (Fig. 4A). Finally, revealing the importance of this residue for STR13/CNX5/
additional modifications occur before the final insertion of MOCS3 function (Matthies et al., 2004).

Moco biosynthesis A Thio-modification of B


cytosolic tRNAs
mitochondrion
CNX2
CNX3
GTP cPMP

tRNA

S uridine
D
cytosol

cPMP S
S CNX5
Cu
CNX7 URM11
SSH
CNX6

MPT
MoO42-
Mg-ATP S
Cu
CNX1
AMP

Moco
2-thiouridine

Fig. 4.  Role of STR13/CNX5 in molybdenum cofactor biosynthesis and thio-modification of cytosolic tRNAs. (A) Simplified representation of molybdenum
cofactor (Moco) biosynthesis in Arabidopsis modified from Kaufholdt et al. (2013). The biosynthesis of Moco starts in the mitochondrial matrix where
guanosine-5′-triphosphate (GTP) is converted to cyclic pyranopterin monophosphate (cPMP) by CNX2 (cofactor for nitrate reductase and xanthine
dehydrogenase 2) and CNX3. The cPMP is then exported to the cytosol and a heteromeric complex, consisting of CNX6 and CNX7, inserts two sulfur
atoms into cPMP and converts it to molybdopterin (MPT). CNX7 receives the sulfur atoms from STR13/CNX5 but the primary sulfur donor of CNX5 (X-S)
is yet unknown. During the last step, one molybdenum atom is inserted via the activity of CNX1, finally leading to the mature Moco. (B) Synthesis of
thiolated nucleoside in tRNAs. STR13/CNX5 provides sulfur to its partner, URM11 (ubiquitin-related modifier 11) and possibly URM12 (not represented),
which is essential for the thio-modification (such as s2U34) of cytosolic tRNAs (Nakai et al., 2012).
Roles of sulfurtransferases in sulfur trafficking in plants  |  4147

Implication in thio-modification of tRNAs to accumulate arsenic, and thus represent good candidates
for arsenic phytoremediation approaches (Peer et  al., 2006).
In addition to its role in Moco biosynthesis, STR13 is also Meanwhile, arsenic accumulation in crops is a challenging issue
involved in the thio-modification of cytosolic tRNAs (Fig. and particularly in rice, which is the staple diet for more than
4B; Nakai et  al., 2012). The thio-modification of tRNAs is half of the world’s population (Williams et al., 2006). Hence,
one out of ca a hundred possible post-transcriptional modi- the investigation of how plants cope with this useless metal is
fications that can influence tRNA folding or stability that a decisive research area and recent advances have been made
would consequently affect translation and ultimately growth in Arabidopsis, implicating two STR isoforms, AtSTR5 and
(Phizicky and Hopper, 2010). The 2-thiouridine s2U corres- AtSTR17a.
ponds to the 2-thio modification that is universally found at According to the involvement of dual-specificity CDC25
U34 of tRNA species. In yeast and higher eukaryotes, U34 in phosphatases in cell cycle regulation, AtSTR5 exhibits phos-
tRNALys (anticodon UUU), tRNAGln (anticodon UUG) and phatase activity in vitro and enhances cyclin-dependent serine/
tRNAGlu (anticodon UUC) are modified and sulfurated to threonine kinase activity of purified CDK complexes via tyro-
form 5-methyl-2-thiouridine derivatives (xm5s2U), the most sine dephosphorylation (Landrieu et  al., 2004a,b). Also, the
important being the 5-methoxycarbonylmethyl-2-thiouridine

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


heterologous expression of AtSTR5 in fission yeast reduces the
(mcm5s2U) modification in eukaryotic cytosolic tRNAs size of mitotic cells and Arabidopsis str5 mutants are hyper-
(Ranjan and Rodnina, 2017). In eukaryotes, the biosynthesis sensitive to a replication blocking agent (Sorrell et  al., 2005;
of s2U in cytosolic tRNAs involves a protein thiocarboxylate Spadafora et al., 2011). In parallel, other studies have suggested
as intermediate sulfur donor that is functionally and evolu- the involvement of AtSTR5 in arsenate reduction. Interestingly,
tionarily related to the ubiquitin-like post-translational modi- both tyrosine phosphatase and arsenate reductase activities
fication system as demonstrated notably in yeast (Nakai et al., depend on the presence of the catalytic cysteine present in
2008). Hence, in yeast, both Urm1 and the yeast STR13 the active site motif (HCX5R) of the Rhd domain (Landrieu
ortholog Uba4 are involved in the thio-modification of cyto- et  al., 2004a). Plant STR5 genes complement the arsenate-
solic tRNAs. Consistently, cytosolic tRNAs from Δurm1 and sensitive phenotype of the E. coli arsC and yeast acr2 mutants
Δuba4 cells are not thiolated (Leidel et al., 2009). Arabidopsis and recombinant proteins display arsenate reductase activity
STR13/CNX5 can fully complement the yeast Δuba4 mu- in vitro (Dhankher et al., 2006; Ellis et al., 2006). Accordingly,
tant. Both cnx5-1 and cnx5-2 mutants exhibit a dwarf pheno- Arabidopsis STR5 down-regulated lines are sensitive to ar-
type with slightly green and morphologically aberrant leaves senate treatments and present a hyperaccumulation of arsenic
and in addition, the thio-modification of cytosolic tRNAs is when they are cultured in vitro for 3 weeks (Dhankher et al.,
impaired (Nakai et  al., 2012). Interestingly, Arabidopsis pos- 2006). Moreover, a 2-fold higher arsenate reductase activity
sesses two orthologs of ScUrm1, URM11 and URM12, both was observed in AtSTR5-overexpressing lines in comparison
of them capable of complementing the yeast Δurm1 mutant with wild-type plants (Dissmeyer et  al., 2009). Puzzlingly, in
(Nakai et  al., 2012). As URM11 is the predominantly ex- another study the content of arsenic species was not modified
pressed isoform, it was suggested that this protein is primarily in both Arabidopsis STR5 knockout or overexpression lines
responsible for the thio-modification of cytosolic tRNAs. exposed to arsenate for 0.5, 2, 6 or 24  h as compared with
Nevertheless, an Arabidopsis urm11-1 null mutant did not wild-type plants (Liu et  al., 2012). Another piece of positive
exhibit any visible phenotypic changes, although the mutant evidence is, however, that transgenic tobacco plants expressing
lacked thio-modified cytosolic tRNAs. Hence, the strong/le- AtSTR5 are more tolerant to arsenic than wild-type plants
thal phenotype observed for str13/cnx5 mutants should be re- (Nahar et al., 2017). The subcellular localization of AtSTR5 is
lated to a defect in Moco biosynthesis, not in cytosolic tRNA also subject to some uncertainty. An experiment using the first
modifications (Nakai et  al., 2012). Altogether, these data re- 100 amino acids of AtSTR5 fused to green fluorescent pro-
veal that STR13 represents an essential link between these two tein (GFP) demonstrated a mitochondrial localization (Narsai
sulfur-dependent biosynthetic pathways but its sulfur donor is et al., 2011). This might correlate with the N-terminal trun-
yet unidentified (Fig. 4). In human cells, a cytosolic form of cation at the position of the Ser20 reported in an N-terminal
the cysteine desulfurase NFS1 is proposed to provide sulfur to proteome study performed with Arabidopsis seedlings (Venne
the MOCS3 ortholog (Marelja et al., 2008, 2013). In plants, the et  al., 2015). Nonetheless, this mitochondrial localization of
sole CD isoform in the cytosol is ABA3. However, aba3 mu- AtSTR5 would not be compatible with its described func-
tants (aba3-1, aba3-2, los5-1, los5-2) have less severe phenotypes tion in cell cycle regulation, which would require a nuclear
(Xiong et al., 2001; Zhong et al., 2010) than cnx5/str13 mutants localization. More experimental evidence is required here
indicating that it should not provide sulfur to STR13, leaving to validate whether AtSTR5 is indeed expressed in multiple
the question of the sulfur source open. subcellular compartments.
The involvement of AtSTR17a in arsenate resistance/ac-
Roles of STR5 and STR17a in arsenate cumulation has been independently identified by two groups
performing genome-wide association studies focusing either
tolerance
on root growth recovery after arsenate exposure in a collection
The presence of inorganic arsenic in soils and fresh water of 82 Arabidopsis accessions (Sánchez-Bermejo et al., 2014) or
is a widely recognized problem. Some plant species, such as on arsenate accumulation in leaves of 349 Arabidopsis acces-
the ferns Pteris vittata and Pityrogamma calomelanos, are known sions (Chao et al., 2014). Based on expression profile analysis,
4148 | Selles et al.

AtSTR17a is highly expressed in roots (Fig. 1). Experiments transferred to GSH. The resulting GSSH could then be oxi-
using AtSTR17a–GFP fusion reporter revealed a nucleo- dized to sulfite (SO32−) by the sulfur dioxygenase ethylmalonic
cytosolic localization (Sánchez-Bermejo et al., 2014) and an ac- encephalopathy protein 1 (ETHE1) (Krüßel et  al., 2014).
cumulation in root hairs, epidermis, and pericycle cells (Chao Sulfite is likely preferentially transferred to peroxisomes and
et al., 2014). As AtSTR17a displays an arsenate reductase ac- oxidized to sulfate by SO (Brychkova et al., 2007, 2013; Lang
tivity and Arabidopsis str17a mutants significantly accumulate et  al., 2007), but it could eventually be further converted to
arsenic in leaves, Chao and co-workers proposed that arsenate thiosulfate (S2O32−) via a persulfidated STR1 (Fig. 5B; Krüßel
reduction by AtSTR17a in roots and notably in the pericycle et al., 2014; Höfler et al., 2016). A role of STR1 in sulfite de-
may limit arsenic loading into the xylem, avoiding its transfer toxification was suggested by the observation of a higher thio-
to the shoots. The difference in the V[G/A]CXXGXR active sulfate level and STR-dependent sulfite-consuming activity in
site motif found in AtSTR17a compared with the canon- RNAi-silenced plants for SO in comparison with wild-type
ical HCX5R active site encountered in yeast and Arabidopsis plants (Brychkova et  al., 2013). Instead of being transferred
ACR2 led to naming AtSTR17a as ARQ1 (arsenate reduc- to GSH, the persulfide group formed on the catalytic Cys of
tase QTL1) or HAC1 (high arsenic content 1)  (Chao et  al., STR1 could be transmitted to other sulfur acceptor molecules

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


2014; Sánchez-Bermejo et al., 2014). Both groups also sought or could lead to the production of hydrogen sulfide (H2S)
to establish a link between AtSTR5 and AtSTR17a in arsenate by reaction with mitochondrial disulfide reductases, notably
tolerance by investigating the impact of AtSTR5/ACR2 mu- TRXo1 and TRXo2 (Fig. 5B). This aspect is further discussed
tation on their respective experimental schemes. No significant in the next section. An alternative route by which STR1 could
changes of the root growth recovery or of arsenate leaf con- obtain sulfur atoms and deliver them to target compounds/
tent were observed in a str5 knockout mutant and the analysis proteins without relying on 3-MP (Fig. 5) is via the mitochon-
of a double str17a str5 mutant did not highlight an epistasic drial cysteine desulfurase NFS1 because an interaction with
interaction between the two genes (Chao et al., 2014; Sánchez- the MST TUM1 has been reported in human and yeast (Noma
Bermejo et  al., 2014). Altogether, these observations support et al., 2009; Fräsdorf et al., 2014).This enzyme catalyses cysteine
a role of both proteins in arsenate tolerance but raise some desulfuration into alanine, providing the sulfur required for de
questions about their respective roles, possibly pointing to their novo Fe–S cluster biogenesis (Couturier et  al., 2013). In this
involvement in distinct mechanisms/pathways. pathway, NFS1 forms an assembly complex with several other
From a molecular point of view, it is informative to correlate proteins and the persulfide group formed on its active site cyst-
the polymorphism present in naturally variable sequences to eine residue provides the sulfur atoms used for building Fe–S
some phenotypes, as was done for AtSTR17a between the clusters. It is worth noting that a similar reaction is catalysed by
Col-0 ecotype and two arsenate hypersensitive ecotypes, Kas-1 NFS2 in chloroplasts (Couturier et  al., 2013). Whether plant
and Kr-0. In Kr-0, the AtSTR17a gene exhibits a premature NFS1 and NFS2 do interact with some STRs is not known.
stop codon leading to the synthesis of an inactive protein de- So far, STR1 is the sole MST identified in Arabidopsis
void of the Rhd domain (Chao et  al., 2014). In Kas-1, the mitochondria (Nakamura et  al., 2000; Papenbrock and
AtSTR17a protein has differences in the N-terminal part but Schmidt, 2000; Senkler et al., 2016). The second MST isoform,
also four internal substitutions in the Rhd domain. A chimeric named STR2 and displaying a high sequence identity with
AtSTR17a Kas-1 protein bearing the N-terminal extension STR1 (79%), was shown to be localized in the cytosol using
and the internal substitutions found in the Col-0 allele is able immunoblots on fractionated cellular extracts and GFP fu-
to complement the arsC mutant. Concomitantly, mutagenesis sion (Nakamura et al., 2000; Papenbrock and Schmidt, 2000).
analysis of the AtSTR17a Col-0 protein bearing a combination However, a longer STR2 transcript (At1g16460.2) is also ex-
of three substitutions from the AtSTR17a Kas-1 protein dem- pressed and encodes a protein with an N-terminal extension
onstrates a non-additive effect of these mutations, but specific of 24 residues that is predicted to constitute a mitochondrial
substitution pairs needed for loss of activity (Sánchez-Bermejo targeting sequence. Hence, this will have to be experimentally
et  al., 2014). Altogether, these data illustrate the natural vari- addressed even though STR2 was not detected in a recent
ability in the AtSTR17a-related arsenate resistance/accumu- mitochondrial proteome analysis, unlike STR1 (Senkler et al.,
lation encountered in different Arabidopsis accessions and the 2016). The study of str mutants indicates that STR1 accounts
complexity of structural features that can promote or desta- for 50–80% of the MST activity whereas str2 mutant seedlings
bilize enzymatic reactions catalysed by STR members. have a wild-type MST activity level (Mao et al., 2011; Höfler
et al., 2016). However, their physiological and molecular roles
in planta have not been elucidated because str2 null mutants do
not have any phenotype and str1 null mutants show a shrunken
Multiple roles for the mitochondrial STR1
seed phenotype with ~87.5% of the embryos arrested at the
In plants, three pathways for cysteine degradation have been heart stage (Mao et al., 2011). The residual 12.5% can develop
described (for a detailed review see Hildebrandt et al., 2015), further and show a normal development of vegetative tissue.
the mitochondrial STR1 contributing to one of these. This Importantly, the double str1 str2 mutant is not viable indicating
pathway starts with the transamination of cysteine to 3-MP, a that both STRs ensure an essential function at least for embryo
reaction catalysed by a so far unknown aminotransferase (Fig. development, but that STR1 likely has a predominant role
5A). STR1 forms pyruvate from 3-MP, the persulfide group (Mao et al., 2011). Similar to STR1, ETHE1 null mutants show
concomitantly formed on the catalytic cysteine being eventually an embryo-lethal phenotype (Krüßel et al., 2014).Whereas this
Roles of sulfurtransferases in sulfur trafficking in plants  |  4149

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


Fig. 5.  Confirmed or hypothetical roles of STR1 in sulfur trafficking in mitochondria. (A) Persulfidation of mitochondrial STR1. This might occur upon
reaction with 3-mercaptopyruvate (3MP) formed from cysteine by a so far unknown cysteine aminotransferase (CAT) or via the transfer of a persulfide
formed on the NFS1 cysteine desulfurase. Whether the latter pathway operates in plants in unknown but the human and yeast STR and NFS orthologs
were shown to interact. (B) Putative reaction pathways of persulfidated STR1. (1) A reaction with reduced glutathione (GSH) generates glutathione
persulfide (GSSH) that can then either react with another GSH molecule leading to the release of hydrogen sulfide (H2S) and oxidized glutathione (GSSG)
or be used as substrate by the sulfur dioxygenase ethylmalonic encephalopathy protein 1 (ETHE1) to form sulfite (SO32−). Sulfite is either detoxified
in sulfate by sulfite oxidase after its transport in peroxisomes, or converted to thiosulfate in mitochondria by persulfidated STR1. (2) A reaction with
thioredoxin o (TRXo) isoforms would lead to H2S formation. (3) In addition to the already described reaction with GSH, TRXo, or sulfite, a reaction with
mitochondrial protein or non-protein acceptors such as cysteine results in the formation persulfidated proteins or of low molecular mass per/polysulfides,
respectively.

would fit with their functional association described in animals (García et al., 2013). In the absence of additional data, for in-
(Kabil and Banerjee, 2014) and with the existence of naturally stance about cyanide, sulfite, and thiosulfate levels in both str1
occurring chimeric proteins containing a sulfur dioxygenase and str2 mutants, a contribution of STR1/2 in cyanide detoxi-
domain fused to one or several Rhd domains (Shen et al., 2015; fication remains uncertain.
Motl et  al., 2017), the mutants have substantial differences.
A  knockdown mutant of ETHE1 (ethe1-1 line) with 1% of
wild-type sulfur dioxygenase activity produces seeds morpho- Production of hydrogen sulfide by STRs
logically indistinguishable from the wild-type although with a
delayed embryo development. Moreover, germination, vegeta- The study of H2S production and its associated physiological
tive growth under short day conditions, and senescence are dif- and signaling roles in plants but also in other organisms is a
ferentially impacted in the ethe1-1 and str1-1 mutants (Krüßel recently emerging area, many aspects remaining mysterious,
et al., 2014; Höfler et al., 2016). notably in plants (Mustafa et al., 2009; Paul and Snyder, 2012;
Hence, it might be that STR1/2 possess additional physio- Greiner et  al., 2013; Longen et  al., 2016; Aroca et  al., 2017).
logical functions in this compartment. Mao and colleagues In plants, H2S is linked to the regulation of many develop-
speculated about a role of both STRs in the protection of mental stages and to stress response mechanisms, being pos-
developing embryos by converting cyanide, a co-product of sibly produced in several subcellular compartments (Hancock
ethylene biosynthesis, to the less toxic thiocyanate (Mao et al., and Whiteman, 2014). Noticeably, it is produced in chloro-
2011) as proposed previously for the mitochondrial MST from plasts in the frame of the reductive assimilation of sulfate
rat to protect cytochrome c oxidase against cyanide (Nagahara (Takahashi et  al., 2011), in mitochondria through the syn-
et al., 1999). However, a direct role in cyanide detoxification thesis of β-cyanoalanine by CAS-C1 (Álvarez et  al., 2012)
seems unlikely because enzymes catalysing this reaction use and in the cytosol through cysteine degradation by L-cysteine
thiosulfate, which is not a good substrate for STR1/2, that ra- desulfhydrase 1 (DES1) (Álvarez et al., 2010). In mammals, H2S
ther produce thiosulfate (Höfler et al., 2016). Hence, an indirect is produced by the side-reactions of cystathionine β-synthase
contribution is possible if STR1/2 produce thiosulfate that is and cystathionine γ-lyase, two enzymes of the trans-sulfuration
used by other mitochondrial STRs, such as STR10. Still, mito- pathway that normally convert cysteine to homocysteine (re-
chondria already contain the β-cyanoalanine synthase CAS- viewed in Kabil and Banerjee, 2014). In a third pathway, H2S
C1 that is involved in the detoxification of cyanide through is released when persulfide-bound forms of the mitochondrial
the formation of β-cyanoalanine (García et  al., 2010). For MST interact with a reductant such as TRX or dihydrolipoic
comparison purpose, it is worth noting that the cas-c1 mutant acid as shown in mice (Shibuya et  al., 2009; Mikami et  al.,
accumulates cyanide notably in roots, strongly inhibiting root 2011;Yadav et al., 2013).The released H2S may serve for modi-
hair elongation and altering the plant response to pathogens fying specific proteins (see below). Interestingly, in mouse
4150 | Selles et al.

hepatocytes lacking both TRX reductase and glutathione re- reactions with some partner proteins (Mishanina et al., 2015).
ductase the level of persulfidated proteins is markedly elevated Accordingly, seven GRX isoforms and 16 TRX isoforms were
(Dóka et al., 2016). Although this underlines the importance retrieved as persulfidated proteins in Arabidopsis (Aroca et al.,
of both reducing systems for protein persulfidation, their exact 2017). Such a relay with TRX or GRX may be convenient to
contribution remains unclear. It may simply be that they serve achieve an additional degree of specificity towards target pro-
for the reduction of persulfidated proteins. teins (Fig. 6).
A connection between STR1 and the reducing systems seems The signaling functions of H2S likely rely on the reactions
true in plants as well given that STR1 has been repeatedly iso- with specific proteins leading to their persulfidation. It is im-
lated by pull-down approaches as a partner of TRX including portant to mention that H2S will react only with oxidized
the mitochondrial TRXo1 isoform from Arabidopsis (Balmer cysteines, possibly sulfenylated, nitrosylated, glutathionylated,
et al., 2004;Yoshida et al., 2013; Pérez-Pérez et al., 2017). In this or forming a disulfide bond. Hence, a complex interplay
species, the STR1–TRXo1 interaction has been confirmed exists between all these redox post-translational modifica-
in vivo using bimolecular fluorescence complementation ex- tions. Alternatively, H2S-mediated protein persulfidation oc-
periments (Henne et  al., 2015). Interestingly, in vivo inter- curs when polysulfides (RS(S)nH and RS(S)nSR) derived from

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


actions between TSTs, such as STR14, STR15, and STR16, H2S react with cysteine thiol or thiolate (Greiner et al., 2013).
and chloroplastic TRXs have been reported suggesting that Remarkable amounts of these polysulfides have been detected
the STR–TRX interaction is not restricted to MSTs. These in vivo in mammalian cells in addition to cysteine-persulfide
STRs may be also involved in H2S biogenesis in addition to (Cys-SSH) and glutathione-persulfide (GSSH) (Ida et  al.,
their capacity to perform trans-persulfidation reactions. An 2014). All persulfide species present in cells, i.e. polysulfides
STR–GRX interaction has not been reported so far but nat- (notably H2S2 and H2S3), low molecular mass persulfides (Cys-
ural fusion proteins containing both protein domains exist in SSH and GSSH) and persulfidated protein cysteine residues,
some prokaryotes. If such an interaction is confirmed, it will be have been designated as forming the bound sulfane sulfur
important to make a distinction between the TRX and GSH/ forms. The levels of total persulfidated species in the brain of
GRX reducing systems (Fig. 6). Indeed, the majority of plant MST-KO mice are less than 50% of that in the brain of wild-
TRXs employ two cysteines to reduce disulfide substrates type mice indicating that mitochondrial MST is indeed an im-
whereas GRXs mostly use a single cysteine (even though portant actor (Kimura et al., 2015, 2017). Similarly, E. coli SseA
there is another cysteine in the active site signature). Hence, is involved in the production of reactive sulfane sulfur and not-
the presence of a persulfide group on a catalytic cysteine of a ably GSSH and GSSSH (Li et al., 2019).
TRX should be short-lived because of the presence of a re- Whether plant MST orthologs, STR1 and STR2, partici-
solving cysteine and should lead to the release of H2S. On the pate in the production of such molecules in plants is poorly
contrary, a persulfide group formed on the catalytic cysteine documented. Nevertheless, in addition to its interaction with
of a GRX may be more stable and allow trans-persulfidation TRXo1 leading to H2S release, Arabidopsis STR1 is able to

Fig. 6.  Possible interaction of STRs with the cellular reducing components. The reaction of persulfidated STRs with dithiol thioredoxin (TRX) or
glutaredoxin (GRX) isoforms generates H2S (1) whereas reaction with monothiol proteins/molecules (monothiol GRX, GSH, cysteine) generates
persulfidated molecules that could participate to subsequent trans-persulfidation reactions (2).
Roles of sulfurtransferases in sulfur trafficking in plants  |  4151

produce GSSH (Höfler et  al., 2016). Moreover, a proteomic residue will represent a way for their post-translational regula-
analysis identified 2015 persulfidated proteins from leaves of tion. Although limited evidence exist so far, several Arabidopsis
wild-type Arabidopsis plants (Aroca et  al., 2017). A  similar STR isoforms were identified as possessing reversibly oxi-
number (2130 persulfidated proteins) and almost the same dized cysteines (Slade et al., 2015) and STR1 was identified as
proteins (85% of concordance) were identified in the des1 mu- sulfenylated in plants treated with H2O2 (Akter et al., 2015; De
tant, which is defective for a cytosolic protein generating H2S Smet et al., 2019). This may further link STRs with the TRX
(Aroca et al., 2017). This suggests that the major factors pro- and GRX reducing systems besides their connection for the
moting protein persulfidation may indeed be STR1 and STR2 formation of H2S and for protein (trans)-persulfidation.
but also other STRs or yet uncharacterized proteins. In mam-
mals, the ubiquitous mitochondrial cysteinyl-tRNA synthetase
has been characterized as the major Cys-SSH synthase in vivo, Acknowledgements
playing an important role in the formation of both low mo- The salary of AM was funded by grants of the French National Research
lecular mass polysulfides and persulfidated proteins (Akaike Agency (ANR) as part of the ‘Investissements d’Avenir’ program (ANR-
et al., 2017). To conclude, elucidating the molecular mechan- 11-LABX-0002-01, Lab of Excellence ARBRE) and of the SULPAR

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


isms underlying H2S-mediated protein persulfidation and its (17_GE2_016) GrandEst region contract. AM also greatly appreciates a
regulation in plants should become a priority and more intense postdoc fellowship from the Alexander von Humboldt Foundation. The
investigations are necessary to understand the effect of these salary of BS and the work on plant sulfurtransferases are supported by
modifications on protein activity and function. grant of the ANR-16-CE20-0012 contract.

References
Conclusions Akaike  T, Ida  T, Wei  FY, et  al. 2017. Cysteinyl-tRNA synthetase gov-
erns cysteine polysulfidation and mitochondrial bioenergetics. Nature
Despite the physiological importance of many sulfur-containing Communications 8, 1177.
molecules, the molecular mechanisms associated with the traf- Akter  S, Huang  J, Bodra  N, et  al. 2015. DYn-2 based identification of
ficking and specific delivery of sulfur atoms to their acceptors Arabidopsis sulfenomes. Molecular & Cellular Proteomics 14, 1183–1200.
remain unclear in a number of pathways. Through their ability Alphey  MS, Williams  RA, Mottram  JC, Coombs  GH, Hunter  WN.
to act as sulfur carrier proteins, STRs fulfill important and di- 2003. The crystal structure of Leishmania major 3-mercaptopyruvate
sulfurtransferase. A  three-domain architecture with a serine protease-
verse roles in these reactions in all organisms and more particu- like triad at the active site. The Journal of Biological Chemistry 278,
larly in plants. Accordingly, the STR family in higher plants is 48219–48227.
composed on average of 20 members that are quite divergent Álvarez  C, Calo  L, Romero  LC, García  I, Gotor  C. 2010. An
considering their primary sequence, protein domain organiza- O-acetylserine(thiol)lyase homolog with L-cysteine desulfhydrase activity
regulates cysteine homeostasis in Arabidopsis. Plant Physiology 152,
tion, and subcellular localization. The best documented func- 656–669.
tions for STR members include roles of STR5 and STR17a Álvarez C, García I, Romero LC, Gotor C. 2012. Mitochondrial sulfide
in arsenate tolerance, of STR13/CNX5 in Moco biosynthesis detoxification requires a functional isoform O-acetylserine(thiol)lyase C in
and thio-modification of cytosolic tRNAs, and of STR1 in Arabidopsis thaliana. Molecular Plant 5, 1217–1226.
cysteine degradation and H2S formation. Aroca  A, Benito  JM, Gotor  C, Romero  LC. 2017. Persulfidation
proteome reveals the regulation of protein function by hydrogen sulfide in
Surprisingly, the exact roles of other STRs and more par- diverse biological processes in Arabidopsis. Journal of Experimental Botany
ticularly those localized in chloroplasts are yet to be explored. 68, 4915–4927.
This is all the more important as these isoforms represent Babicki  S, Arndt  D, Marcu  A, Liang  Y, Grant  JR, Maciejewski  A,
nearly half of all STRs and sulfide and sulfite are produced in Wishart  DS. 2016. Heatmapper: web-enabled heat mapping for all.
Nucleic Acids Research 44, W147–W153.
plastids during sulfate reduction. STR4/thylakoid rhodanese-
Balmer Y, Vensel WH, Tanaka CK, et al. 2004. Thioredoxin links redox to
like protein (TROL), which is inactive as STR, is required for the regulation of fundamental processes of plant mitochondria. Proceedings
tethering ferredoxin-NADP oxidoreductase to the thylakoid of the National Academy of Sciences, USA 101, 2642–2647.
membrane (Jurić et al., 2009). This interaction is proposed to Bartels A. 2006. Functional characterisation of sulfurtransferase proteins in
control the flow of photosynthetic electrons prior to activation higher plants. PhD thesis, Leibniz Universität Hannover.
of pseudo-cyclic electron flow (Vojta et al., 2015). Bartels  A, Mock  HP, Papenbrock  J. 2007. Differential expression of
Arabidopsis sulfurtransferases under various growth conditions. Plant
In addition to genetic and physiological studies, experi- Physiology and Biochemistry 45, 178–187.
ments aiming at identifying STR partners are required and Bauer  M, Dietrich  C, Nowak  K, Sierralta  WD, Papenbrock  J. 2004.
should help in defining their precise roles and eventually lead Intracellular localization of Arabidopsis sulfurtransferases. Plant Physiology
to the identification of new STR-dependent pathways. In vitro 135, 916–926.
analyses of the interactions between STRs and sulfur donors/ Behshad E, Bollinger JM Jr. 2009. Kinetic analysis of cysteine desulfurase
CD0387 from Synechocystis sp. PCC 6803: formation of the persulfide
acceptors appear crucial and this necessitates more systematic intermediate. Biochemistry 48, 12014–12023.
exploration of the biochemical and structural properties of Bernard DG, Cheng Y, Zhao Y, Balk J. 2009. An allelic mutant series of
STR isoforms alone or upon interaction. ATM3 reveals its key role in the biogenesis of cytosolic iron-sulfur proteins in
Whether and how STR activity is regulated is another as- Arabidopsis. Plant Physiology 151, 590–602.
pect to investigate in the future. As the role of STRs in sulfur Bittner F, Oreb M, Mendel RR. 2001. ABA3 is a molybdenum cofactor
sulfurase required for activation of aldehyde oxidase and xanthine dehydro-
exchange reactions relies on the presence of a catalytic cyst- genase in Arabidopsis thaliana. The Journal of Biological Chemistry 276,
eine in their Rhd domain, any modifications of this mandatory 40381–40384.
4152 | Selles et al.
Blumenthal KM, Heinrikson RL. 1971. Structural studies of bovine liver Hancock  JT, Whiteman  M. 2014. Hydrogen sulfide and cell signaling:
rhodanese. I. Isolation and characterization of two active forms of the en- team player or referee? Plant Physiology and Biochemistry 78, 37–42.
zyme. The Journal of Biological Chemistry 246, 2430–2437. Hatzfeld  Y, Saito  K. 2000. Evidence for the existence of rhodanese
Bordo D, Bork P. 2002. The rhodanese/Cdc25 phosphatase superfamily. (thiosulfate:cyanide sulfurtransferase) in plants: preliminary characteriza-
Sequence-structure-function relations. EMBO Reports 3, 741–746. tion of two rhodanese cDNAs from Arabidopsis thaliana. FEBS Letters 470,
Brychkova G, Grishkevich V, Fluhr R, Sagi M. 2013. An essential role for 147–150.
tomato sulfite oxidase and enzymes of the sulfite network in maintaining leaf Henne  M, König  N, Triulzi  T, Baroni  S, Forlani  F, Scheibe  R,
sulfite homeostasis. Plant Physiology 161, 148–164. Papenbrock J. 2015. Sulfurtransferase and thioredoxin specifically interact
Brychkova G, Xia Z, Yang G, Yesbergenova Z, Zhang Z, Davydov O, as demonstrated by bimolecular fluorescence complementation analysis
Fluhr  R, Sagi  M. 2007. Sulfite oxidase protects plants against sulfur di- and biochemical tests. FEBS Open Bio 5, 832–843.
oxide toxicity. The Plant Journal 50, 696–709. Hildebrandt TM, Grieshaber MK. 2008. Redox regulation of mitochon-
Burow  M, Kessler  D, Papenbrock  J. 2002. Enzymatic activity of the drial sulfide oxidation in the lugworm, Arenicola marina. The Journal of
Arabidopsis sulfurtransferase resides in the C-terminal domain but is Experimental Biology 211, 2617–2623.
boosted by the N-terminal domain and the linker peptide in the full-length Hildebrandt  TM, Nunes  Nesi  A, Araújo  WL, Braun  HP. 2015. Amino
enzyme. Biological Chemistry 383, 1363–1372. acid catabolism in plants. Molecular Plant 8, 1563–1579.
Chao  DY, Chen  Y, Chen  J, Shi  S, Chen  Z, Wang  C, Danku  JM, Höfler  S, Lorenz  C, Busch  T, Brinkkötter  M, Tohge  T, Fernie  AR,
Zhao FJ, Salt DE. 2014. Genome-wide association mapping identifies a Braun  HP, Hildebrandt  TM. 2016. Dealing with the sulfur part of cyst-

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


new arsenate reductase enzyme critical for limiting arsenic accumulation in eine: four enzymatic steps degrade l-cysteine to pyruvate and thiosulfate in
plants. PLoS Biology 12, e1002009. Arabidopsis mitochondria. Physiologia Plantarum 157, 352–366.
Cipollone R, Ascenzi P, Visca P. 2007. Common themes and variations Huang  GT, Yu  JS. 2016. Enzyme catalysis that paves the way for
in the rhodanese superfamily. IUBMB Life 59, 51–59. S-sulfhydration via sulfur atom transfer. The Journal of Physical Chemistry.
Colnaghi  R, Cassinelli  G, Drummond  M, Forlani  F, Pagani  S. 2001. B 120, 4608–4615.
Properties of the Escherichia coli rhodanese-like protein SseA: contribution Ida  T, Sawa  T, Ihara  H, et  al. 2014. Reactive cysteine persulfides and
of the active-site residue Ser240 to sulfur donor recognition. FEBS Letters S-polythiolation regulate oxidative stress and redox signaling. Proceedings
500, 153–156. of the National Academy of Sciences, USA 111, 7606–7611.
Cornilescu  G, Vinarov  DA, Tyler  EM, Markley  JL, Cornilescu  CC. Jurić  S, Hazler-Pilepić  K, Tomasić  A, et  al. 2009. Tethering of
2006. Solution structure of a single-domain thiosulfate sulfurtransferase ferredoxin:NADP+ oxidoreductase to thylakoid membranes is mediated by
from Arabidopsis thaliana. Protein Science 15, 2836–2841. novel chloroplast protein TROL. The Plant Journal 60, 783–794.
Couturier J, Jacquot JP, Rouhier N. 2013. Toward a refined classification Kabil  O, Banerjee  R. 2014. Enzymology of H2S biogenesis, decay and
of class I dithiol glutaredoxins from poplar: biochemical basis for the defin- signaling. Antioxidants & Redox Signaling 20, 770–782.
ition of two subclasses. Frontiers in Plant Science 4, 518. Kaufholdt  D, Gehl  C, Geisler  M, Jeske  O, Voedisch  S, Ratke  C,
De  Smet  B, Willems  P, Fernandez-Fernandez  AD, Alseekh  S, Bollhöner  B, Mendel  RR, Hänsch  R. 2013. Visualization and quantifi-
Fernie  AR, Messens  J, Van  Breusegem  F. 2019. In vivo detec- cation of protein interactions in the biosynthetic pathway of molybdenum
tion of protein cysteine sulfenylation in plastids. The Plant Journal 97, cofactor in Arabidopsis thaliana. Journal of Experimental Botany 64,
765–778. 2005–2016.
Dhankher  OP, Rosen  BP, McKinney  EC, Meagher  RB. 2006. Kimura  Y, Koike  S, Shibuya  N, Lefer  D, Ogasawara  Y, Kimura  H.
Hyperaccumulation of arsenic in the shoots of Arabidopsis silenced for 2017. 3-Mercaptopyruvate sulfurtransferase produces potential redox regu-
arsenate reductase (ACR2). Proceedings of the National Academy of lators cysteine- and glutathione-persulfide (Cys-SSH and GSSH) together
Sciences, USA 103, 5413–5418. with signaling molecules H2S2, H2S3 and H2S. Scientific Reports 7, 10459.
Dissmeyer N, Weimer AK, Pusch S, et al. 2009. Control of cell prolifer- Kimura  Y, Toyofuku  Y, Koike  S, Shibuya  N, Nagahara  N, Lefer  D,
ation, organ growth, and DNA damage response operate independently of Ogasawara Y, Kimura H. 2015. Identification of H2S3 and H2S produced
dephosphorylation of the Arabidopsis Cdk1 homolog CDKA;1. The Plant by 3-mercaptopyruvate sulfurtransferase in the brain. Scientific Reports 5,
Cell 21, 3641–3654. 14774.
Dóka É, Pader I, Bíró A, et al. 2016. A novel persulfide detection method Klepikova AV, Kasianov AS, Gerasimov ES, Logacheva MD, Penin AA.
reveals protein persulfide- and polysulfide-reducing functions of thioredoxin 2016. A high resolution map of the Arabidopsis thaliana developmental tran-
and glutathione systems. Science Advances 2, e1500968. scriptome based on RNA-seq profiling. The Plant Journal 88, 1058–1070.
Ellis  DR, Gumaelius  L, Indriolo  E, Pickering  IJ, Banks  JA, Salt  DE. Krepinsky K, Leimkühler S. 2007. Site-directed mutagenesis of the active
2006. A novel arsenate reductase from the arsenic hyperaccumulating fern site loop of the rhodanese-like domain of the human molybdopterin syn-
Pteris vittata. Plant Physiology 141, 1544–1554. thase sulfurase MOCS3. Major differences in substrate specificity between
Fräsdorf  B, Radon  C, Leimkühler  S. 2014. Characterization and inter- eukaryotic and bacterial homologs. The FEBS Journal 274, 2778–2787.
action studies of two isoforms of the dual localized 3-mercaptopyruvate Kruse I, Maclean AE, Hill L, Balk J. 2018. Genetic dissection of cyclic
sulfurtransferase TUM1 from humans. The Journal of Biological Chemistry pyranopterin monophosphate biosynthesis in plant mitochondria. The
289, 34543–34556. Biochemical Journal 475, 495–509.
Friso  G, Giacomelli  L, Ytterberg  AJ, Peltier  JB, Rudella  A, Sun  Q, Krüßel  L, Junemann  J, Wirtz  M, et  al. 2014. The mitochondrial sulfur
Wijk  KJ. 2004. In-depth analysis of the thylakoid membrane proteome dioxygenase ETHYLMALONIC ENCEPHALOPATHY PROTEIN1 is required
of Arabidopsis thaliana chloroplasts: new proteins, new functions, and a for amino acid catabolism during carbohydrate starvation and embryo de-
plastid proteome database. The Plant Cell 16, 478–499. velopment in Arabidopsis. Plant Physiology 165, 92–104.
García I, Castellano JM, Vioque B, Solano R, Gotor C, Romero LC. Landrieu  I, da  Costa  M, De  Veylder  L, et  al. 2004a. A small CDC25
2010. Mitochondrial β-cyanoalanine synthase is essential for root hair for- dual-specificity tyrosine-phosphatase isoform in Arabidopsis thaliana.
mation in Arabidopsis thaliana. The Plant Cell 22, 3268–3279. Proceedings of the National Academy of Sciences, USA 101, 13380–13385.
García I, Rosas T, Bejarano ER, Gotor C, Romero LC. 2013. Transient Landrieu I, Hassan S, Sauty M, Dewitte F, Wieruszeski JM, Inzé D,
transcriptional regulation of the CYS-C1 gene and cyanide accumulation De  Veylder  L, Lippens  G. 2004b. Characterization of the Arabidopsis
upon pathogen infection in the plant immune response. Plant Physiology thaliana Arath;CDC25 dual-specificity tyrosine phosphatase. Biochemical
162, 2015–2027. and Biophysical Research Communications 322, 734–739.
Greiner  R, Pálinkás  Z, Bäsell  K, Becher  D, Antelmann  H, Nagy  P, Lang  C, Popko  J, Wirtz  M, Hell  R, Herschbach  C, Kreuzwieser  J,
Dick TP. 2013. Polysulfides link H2S to protein thiol oxidation. Antioxidants Rennenberg  H, Mendel  RR, Hänsch  R. 2007. Sulphite oxidase as
& Redox Signaling 19, 1749–1765. key enzyme for protecting plants against sulphur dioxide. Plant, Cell &
Guretzki S, Papenbrock J. 2011. Characterization of the sulfurtransferase Environment 30, 447–455.
family from Oryza sativa L. Plant Physiology and Biochemistry 49, Lanz ND, Booker SJ. 2015. Auxiliary iron-sulfur cofactors in radical SAM
1064–1070. enzymes. Biochimica et Biophysica Acta 1853, 1316–1334.
Roles of sulfurtransferases in sulfur trafficking in plants  |  4153
Lec  J-C, Boutserin  S, Mazon  H, Mulliert  G, Boschi-Muller  S, Nahar N, Rahman A, Ghosh S, Nawani N, Mandal A. 2017. Functional
Talfournier F. 2018. Unraveling the mechanism of cysteine persulfide for- studies of AtACR2 gene putatively involved in accumulation, reduction and/
mation catalyzed by 3-mercaptopyruvate sulfurtransferases. ACS Catalysis or sequestration of arsenic species in plants. Biologia 72, 520–526.
8, 2049–2059. Nakai  Y, Harada  A, Hashiguchi  Y, Nakai  M, Hayashi  H. 2012.
Leidel  S, Pedrioli  PG, Bucher  T, Brost  R, Costanzo  M, Schmidt  A, Arabidopsis molybdopterin biosynthesis protein Cnx5 collaborates with the
Aebersold  R, Boone  C, Hofmann  K, Peter  M. 2009. Ubiquitin-related ubiquitin-like protein Urm11 in the thio-modification of tRNA. The Journal of
modifier Urm1 acts as a sulphur carrier in thiolation of eukaryotic transfer Biological Chemistry 287, 30874–30884.
RNA. Nature 458, 228–232. Nakai Y, Nakai M, Hayashi H. 2008. Thio-modification of yeast cytosolic
Li K, Xin Y, Xuan G, Zhao R, Liu H, Xia Y, Xun L. 2019. Escherichia coli tRNA requires a ubiquitin-related system that resembles bacterial sulfur
uses separate enzymes to produce H2S and reactive sulfane sulfur from transfer systems. The Journal of Biological Chemistry 283, 27469–27476.
L-cysteine. Frontiers in Microbiology 10, 298. Nakamura  T, Yamaguchi  Y, Sano  H. 2000. Plant mercaptopyruvate
Libiad  M, Motl  N, Akey  DL, Sakamoto  N, Fearon  ER, Smith  JL, sulfurtransferases: molecular cloning, subcellular localization and enzymatic
Banerjee  R. 2018. Thiosulfate sulfurtransferase-like domain-containing 1 activities. European Journal of Biochemistry 267, 5621–5630.
protein interacts with thioredoxin. The Journal of Biological Chemistry 293, Nambi  S, Long  JE, Mishra  BB, Baker  R, Murphy  KC, Olive  AJ,
2675–2686. Nguyen  HP, Shaffer  SA, Sassetti  CM. 2015. The oxidative stress net-
Liu  W, Schat  H, Bliek  M, Chen  Y, McGrath  SP, George  G, Salt  DE, work of Mycobacterium tuberculosis reveals coordination between radical
Zhao FJ. 2012. Knocking out ACR2 does not affect arsenic redox status in detoxification systems. Cell Host & Microbe 17, 829–837.
Arabidopsis thaliana: implications for as detoxification and accumulation in

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


Narsai R, Law SR, Carrie C, Xu L, Whelan J. 2011. In-depth temporal
plants. PLoS ONE 7, e42408. transcriptome profiling reveals a crucial developmental switch with roles for
Longen S, Richter F, Köhler Y, Wittig I, Beck KF, Pfeilschifter J. 2016. RNA processing and organelle metabolism that are essential for germination
Quantitative persulfide site identification (qPerS-SID) reveals protein targets in Arabidopsis. Plant Physiology 157, 1342–1362.
of H2S releasing donors in mammalian cells. Scientific Reports 6, 29808. Noma  A, Sakaguchi  Y, Suzuki  T. 2009. Mechanistic characterization
Luo GX, Horowitz PM. 1994. The sulfurtransferase activity and structure of the sulfur-relay system for eukaryotic 2-thiouridine biogenesis at tRNA
of rhodanese are affected by site-directed replacement of Arg-186 or Lys- wobble positions. Nucleic Acids Research 37, 1335–1352.
249. The Journal of Biological Chemistry 269, 8220–8225.
Nomura H, Komori T, Kobori M, Nakahira Y, Shiina T. 2008. Evidence
Mao  G, Wang  R, Guan  Y, Liu  Y, Zhang  S. 2011. Sulfurtransferases 1 for chloroplast control of external Ca2+-induced cytosolic Ca2+ transients
and 2 play essential roles in embryo and seed development in Arabidopsis and stomatal closure. The Plant Journal 53, 988–998.
thaliana. The Journal of Biological Chemistry 286, 7548–7557.
Pagani S, Forlani F, Carpen A, Bordo D, Colnaghi R. 2000. Mutagenic
Marelja  Z, Mullick  Chowdhury  M, Dosche  C, Hille  C, Baumann  O, analysis of Thr-232 in rhodanese from Azotobacter vinelandii highlighted the
Löhmannsröben  HG, Leimkühler  S. 2013. The L-cysteine desulfurase differences of this prokaryotic enzyme from the known sulfurtransferases.
NFS1 is localized in the cytosol where it provides the sulfur for molybdenum FEBS Letters 472, 307–311.
cofactor biosynthesis in humans. PLoS ONE 8, e60869.
Pantoja-Uceda D, López-Méndez B, Koshiba S, et al. 2005. Solution
Marelja Z, Stöcklein W, Nimtz M, Leimkühler S. 2008. A novel role for structure of the rhodanese homology domain At4g01050(175-295) from
human Nfs1 in the cytoplasm: Nfs1 acts as a sulfur donor for MOCS3, Arabidopsis thaliana. Protein Science 14, 224–230.
a protein involved in molybdenum cofactor biosynthesis. The Journal of
Papenbrock  J, Guretzki  S, Henne  M. 2011. Latest news about the
Biological Chemistry 283, 25178–25185.
sulfurtransferase protein family of higher plants. Amino Acids 41, 43–57.
Matthies A, Rajagopalan KV, Mendel RR, Leimkühler S. 2004. Evidence
Papenbrock J, Schmidt A. 2000. Characterization of two sulfurtransferase
for the physiological role of a rhodanese-like protein for the biosynthesis of
isozymes from Arabidopsis thaliana. European Journal of Biochemistry 267,
the molybdenum cofactor in humans. Proceedings of the National Academy
5571–5579.
of Sciences, USA 101, 5946–5951.
Paul  BD, Snyder  SH. 2012. H2S signalling through protein sulfhydration
Mikami  Y, Shibuya  N, Kimura  Y, Nagahara  N, Ogasawara  Y,
and beyond. Nature Reviews. Molecular Cell Biology 13, 499–507.
Kimura  H. 2011. Thioredoxin and dihydrolipoic acid are required for
3-mercaptopyruvate sulfurtransferase to produce hydrogen sulfide. The Peer  WA, Baxter  IR, Richards  EL, Freeman  JL, Murphy  AS.
Biochemical Journal 439, 479–485. 2006. Phytoremediation and hyperaccumulator plants. In: Tamas  MJ,
Martinoia  E, eds. Topics in current genetics. Molecular biology of metal
Mishanina  TV, Libiad  M, Banerjee  R. 2015. Biogenesis of reactive
homeostasis and detoxification: from microbes to man. Berlin, Heidelberg:
sulfur species for signaling by hydrogen sulfide oxidation pathways. Nature
Chemical Biology 11, 457–464. Springer, 299–340.
Moseler A, Selles B, Rouhier N, Couturier J. 2019. Novel insights into Peltier JB, Ytterberg AJ, Sun Q, van Wijk KJ. 2004. New functions of
the diversity of the sulfurtransferase family in photosynthetic organisms with the thylakoid membrane proteome of Arabidopsis thaliana revealed by a
emphasis on oak. New Phytologist, doi: 10.1111/nph.15870. simple, fast, and versatile fractionation strategy. The Journal of Biological
Chemistry 279, 49367–49383.
Motl  N, Skiba  MA, Kabil  O, Smith  JL, Banerjee  R. 2017. Structural
and biochemical analyses indicate that a bacterial persulfide dioxygenase- Pérez-Pérez  ME, Mauriès  A, Maes  A, Tourasse  NJ, Hamon  M,
rhodanese fusion protein functions in sulfur assimilation. The Journal of Lemaire  SD, Marchand  CH. 2017. The deep thioredoxome in
Biological Chemistry 292, 14026–14038. Chlamydomonas reinhardtii: new insights into redox regulation. Molecular
Plant 10, 1107–1125.
Mueller EG. 2006. Trafficking in persulfides: delivering sulfur in biosynthetic
pathways. Nature Chemical Biology 2, 185–194. Phizicky EM, Hopper AK. 2010. tRNA biology charges to the front. Genes
& Development 24, 1832–1860.
Mustafa AK, Gadalla MM, Sen N, Kim S, Mu W, Gazi SK, Barrow RK,
Yang  G, Wang  R, Snyder  SH. 2009. H2S signals through protein Ploegman JH, Drent G, Kalk KH, Hol WG. 1978. Structure of bovine liver
S-sulfhydration. Science Signaling 2, ra72. rhodanese. I. Structure determination at 2.5 Å resolution and a comparison
of the conformation and sequence of its two domains. Journal of Molecular
Nagahara N, Ito T, Minami M. 1999. Mercaptopyruvate sulfurtransferase Biology 123, 557–594.
as a defense against cyanide toxication: molecular properties and mode of
detoxification. Histology and Histopathology 14, 1277–1286. Ranjan N, Rodnina MV. 2017. Thio-modification of tRNA at the Wobble
position as regulator of the kinetics of decoding and translocation on the
Nagahara  N, Nishino  T. 1996. Role of amino acid residues in the ac- ribosome. Journal of the American Chemical Society 139, 5857–5864.
tive site of rat liver mercaptopyruvate sulfurtransferase. CDNA cloning,
overexpression, and site-directed mutagenesis. The Journal of Biological Sánchez-Bermejo E, Castrillo G, del Llano B, et al. 2014. Natural vari-
Chemistry 271, 27395–27401. ation in arsenate tolerance identifies an arsenate reductase in Arabidopsis
thaliana. Nature Communications 5, 4617.
Nagahara  N, Okazaki  T, Nishino  T. 1995. Cytosolic mercaptopyruvate
sulfurtransferase is evolutionarily related to mitochondrial rhodanese. Schwarz  G, Mendel  RR. 2006. Molybdenum cofactor biosynthesis and
Striking similarity in active site amino acid sequence and the increase in the molybdenum enzymes. Annual Review of Plant Biology 57, 623–647.
mercaptopyruvate sulfurtransferase activity of rhodanese by site-directed Senkler  J, Senkler  M, Eubel  H, Hildebrandt  T, Lengwenus  C,
mutagenesis. The Journal of Biological Chemistry 270, 16230–16235. Schertl P, Schwarzländer M, Wagner S, Wittig I, Braun H-P. 2016. The
4154 | Selles et al.
mitochondrial complexome of Arabidopsis thaliana. The Plant Journal 89, Venne  AS, Solari  FA, Faden  F, Paretti  T, Dissmeyer  N, Zahedi  RP.
1079–1092. 2015. An improved workflow for quantitative N-terminal charge-based frac-
Shen  J, Keithly  ME, Armstrong  RN, Higgins  KA, Edmonds  KA, tional diagonal chromatography (ChaFRADIC) to study proteolytic events in
Giedroc  DP. 2015. Staphylococcus aureus CstB is a novel multidomain Arabidopsis thaliana. Proteomics 15, 2458–2469.
persulfide dioxygenase-sulfurtransferase involved in hydrogen sulfide de- Vojta L, Carić D, Cesar V, Antunović Dunić J, Lepeduš H, Kveder M,
toxification. Biochemistry 54, 4542–4554. Fulgosi  H. 2015. TROL-FNR interaction reveals alternative pathways of
Shibuya N, Tanaka M, Yoshida M, Ogasawara Y, Togawa T, Ishii K, electron partitioning in photosynthesis. Scientific Reports 5, 10085.
Kimura H. 2009. 3-Mercaptopyruvate sulfurtransferase produces hydrogen Weinl S, Held K, Schlücking K, Steinhorst L, Kuhlgert S, Hippler M,
sulfide and bound sulfane sulfur in the brain. Antioxidants & Redox Signaling Kudla  J. 2008. A plastid protein crucial for Ca2+-regulated stomatal re-
11, 703–714. sponses. New Phytologist 179, 675–686.
Slade  WO, Werth  EG, McConnell  EW, Alvarez  S, Hicks  LM. 2015. Williams PN, Islam MR, Adomako EE, Raab A, Hossain SA, Zhu YG,
Quantifying reversible oxidation of protein thiols in photosynthetic or- Feldmann J, Meharg AA. 2006. Increase in rice grain arsenic for regions
ganisms. Journal of the American Society for Mass Spectrometry 26, of Bangladesh irrigating paddies with elevated arsenic in groundwaters.
631–640. Environmental Science & Technology 40, 4903–4908.
Sorrell DA, Chrimes D, Dickinson JR, Rogers HJ, Francis D. 2005. Winter D, Vinegar B, Nahal H, Ammar R, Wilson GV, Provart NJ. 2007.
The Arabidopsis CDC25 induces a short cell length when overexpressed An “Electronic Fluorescent Pictograph” browser for exploring and analyzing
in fission yeast: evidence for cell cycle function. New Phytologist 165, large-scale biological data sets. PLoS ONE 2, e718.
425–428.

Downloaded from https://academic.oup.com/jxb/article/70/16/4139/5485845 by guest on 08 February 2023


Wollers S, Heidenreich T, Zarepour M, Zachmann D, Kraft C, Zhao Y,
Spadafora  ND, Doonan  JH, Herbert  RJ, Bitonti  MB, Wallace  E, Mendel RR, Bittner F. 2008. Binding of sulfurated molybdenum cofactor
Rogers  HJ, Francis  D. 2011. Arabidopsis T-DNA insertional lines for to the C-terminal domain of ABA3 from Arabidopsis thaliana provides in-
CDC25 are hypersensitive to hydroxyurea but not to zeocin or salt stress. sight into the mechanism of molybdenum cofactor sulfuration. The Journal
Annals of Botany 107, 1183–1192. of Biological Chemistry 283, 9642–9650.
Su T, Xu J, Li Y, Lei L, Zhao L, Yang H, Feng J, Liu G, Ren D. 2011. Xiong L, Ishitani M, Lee H, Zhu JK. 2001. The Arabidopsis LOS5/ABA3
Glutathione-indole-3-acetonitrile is required for camalexin biosynthesis in locus encodes a molybdenum cofactor sulfurase and modulates cold
Arabidopsis thaliana. The Plant Cell 23, 364–380. stress- and osmotic stress-responsive gene expression. The Plant Cell 13,
Takahashi H, Kopriva S, Giordano M, Saito K, Hell R. 2011. Sulfur as- 2063–2083.
similation in photosynthetic organisms: molecular functions and regulations Yadav PK, Yamada K, Chiku T, Koutmos M, Banerjee R. 2013. Structure
of transporters and assimilatory enzymes. Annual Review of Plant Biology and kinetic analysis of H2S production by human mercaptopyruvate
62, 157–184. sulfurtransferase. The Journal of Biological Chemistry 288, 20002–20013.
Teschner  J, Lachmann  N, Schulze  J, Geisler  M, Selbach  K, Yoshida K, Noguchi K, Motohashi K, Hisabori T. 2013. Systematic ex-
Santamaria-Araujo J, Balk J, Mendel RR, Bittner F. 2010. A novel role ploration of thioredoxin target proteins in plant mitochondria. Plant & Cell
for Arabidopsis mitochondrial ABC transporter ATM3 in molybdenum co- Physiology 54, 875–892.
factor biosynthesis. The Plant Cell 22, 468–480. Zheng  L, White  RH, Cash  VL, Jack  RF, Dean  DR. 1993. Cysteine
Tomizioli M, Lazar C, Brugière S, et al. 2014. Deciphering thylakoid sub- desulfurase activity indicates a role for NIFS in metallocluster biosynthesis.
compartments using a mass spectrometry-based approach. Molecular & Proceedings of the National Academy of Sciences, USA 90, 2754–2758.
Cellular Proteomics 13, 2147–2167. Zhong R, Thompson J, Ottesen E, Lamppa GK. 2010. A forward gen-
Vainonen JP, Sakuragi Y, Stael S, et al. 2008. Light regulation of CaS, etic screen to explore chloroplast protein import in vivo identifies Moco
a novel phosphoprotein in the thylakoid membrane of Arabidopsis thaliana. sulfurase, pivotal for ABA and IAA biosynthesis and purine turnover. The
The FEBS Journal 275, 1767–1777. Plant Journal 63, 44–59.

You might also like